Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4073 2023-02-17 13:15:23 |
2 format correct + 1 word(s) 4074 2023-02-20 03:31:50 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ojo, O.E.; Kreuzer-Redmer, S. The Role of MicroRNAs in Ruminants. Encyclopedia. Available online: https://encyclopedia.pub/entry/41358 (accessed on 06 September 2024).
Ojo OE, Kreuzer-Redmer S. The Role of MicroRNAs in Ruminants. Encyclopedia. Available at: https://encyclopedia.pub/entry/41358. Accessed September 06, 2024.
Ojo, Oyindamola Esther, Susanne Kreuzer-Redmer. "The Role of MicroRNAs in Ruminants" Encyclopedia, https://encyclopedia.pub/entry/41358 (accessed September 06, 2024).
Ojo, O.E., & Kreuzer-Redmer, S. (2023, February 17). The Role of MicroRNAs in Ruminants. In Encyclopedia. https://encyclopedia.pub/entry/41358
Ojo, Oyindamola Esther and Susanne Kreuzer-Redmer. "The Role of MicroRNAs in Ruminants." Encyclopedia. Web. 17 February, 2023.
The Role of MicroRNAs in Ruminants
Edit

One of the most prosperous subspecies of terrestrial mammals is the ruminant family. They live in a wide range of diverse environments around the world and have a big impact on various ecosystems, industries including agriculture, leisure activities, and cultures. MicroRNAs perform essential functions in a variety of biological processes, including differentiation, cell proliferation, metabolism, development, and inflammation. Numerous studies have demonstrated a connection between circulating microRNAs and ruminant diseases such as mastitis, tuberculosis, foot-and-mouth disease, fasciolosis, and metabolic disorders.

microRNAs ruminant nutrition ruminant physiology

1. Introduction

Feed regurgitation, chewing, salivation, and swallowing are described as rumination. Rumination reduces particle size, increasing microbial activity and making it easier for the digest to pass through the digestive tract. Ruminants digest their meals differently from non-ruminants due to their distinct gastrointestinal systems. Ruminating animals have several physiological traits that assist them in surviving in harsh environmental conditions. As a result, ruminants require different antibiotic-like compounds (mostly ionophores, lipophilic compounds) and direct-fed microbials than non-ruminants. These ruminant-specific products are thought to improve ruminal development, lactation, and production by influencing the generation of short chain fatty acids (SCFA) and microbial protein in the rumen to increase the provision of metabolizable energy and a balanced mix of amino acids [1]. The idea of nutritional value combines knowledge of the nutrients present and their availability, with considerations of the typical consumption rates, flavor, and the impact of the meal on the health of animals and the standard of animal products. Understanding ruminant responses to dietary components and other environmental factors elucidates the significance of the environment–gene expression interaction, including post transcriptional regulation through microRNAs (miRNAs). A better understanding of this interaction presents the opportunity to modulate metabolism, health, and diseases based on nutritional strategies.
Feeds are known to provide a conditioning environment that shapes the genome’s activity and the body’s physiology [2]. Direct studies of feeding methods associated with production characteristics have generally dominated the traditional approach in ruminant nutrition. However, this approach was unable to offer sufficient information regarding nutrient dynamics in the gastrointestinal tract (GIT), their impact on tissue, and molecular mechanisms of metabolism. As a result of advances in ruminant nutrition and physiology, molecular biology, high-throughput technologies, and bioinformatics databases, other approaches, such as nutrigenomics, epigenetics, metagenomics, metabolomics, transcriptomics, and proteomics have become more prominent. The environment modulates gene expression through modifications. One type of that modification of gene expression is epigenetics, and if epigenetics is thought of as a harmonizing process, many phenotypic differences can be simply explained [3]. Another mode of regulation of gene expression, influenceable through environmental factors, is facilitated post-transcriptionally through miRNAs.
Understanding gene regulation, including the effect of nutrition in altering noncoding regulatory RNA such as miRNAs, is one of the foundations of the biological and molecular underpinnings of how diet affects animal nutrition and phenotypic variation [4]. MiRNAs are single-stranded noncoding RNA molecules, which are 18–25 nucleotides (nts) long, that contribute to posttranscriptional gene regulation by binding (usually with imperfect complementarity) to the 3-untranslated (UTR) of a target messenger RNA (mRNA), resulting in translation degradation or inhibition [5]. Each miRNA is considered to have several targets, and multiple miRNAs can converge on a single mRNA, implying that these fascinating molecules play a significant regulatory role in metabolism and development [6]. Lin-4 and lethal-7 (let-7) were the first miRNAs discovered, and they were determined to be critical for developmental timing in Caenorhabditis elegans [7]. Following that, a mammalian let-7 family with twelve members expressed from eight separate loci (let-7a-1, -2, -3; let-7b; let-7c; let-7d; let-7e; let-7f-1, -2; let-7g; let-7i; miR-98) was discovered and characterized [8]. Even though they are found all over the genome, many let-7 miRNA family members are coordinated during development, and their regulation has been found to involve RNA-binding proteins, such as Lin28, which inhibits let-7 biogenesis. Since the discovery of the first miRNA (lin-4) in 1993 [9], advancements in next-generation sequencing (NGS, also referred as second-generation sequencing) and third-generation sequencing (TGS) technology have ushered in a new age and the ability to rapidly detect numerous classes of small RNA molecules, including miRNA, in various biological samples [10]. Repeated free-thaw cycles and long-term storage have been found to be stable for miRNAs in many biological samples, which make miRNAs a robust biological marker [11]. Various pathways, such as direct fusion, internalization, and receptor-mediated interactions, are hypothesized to be involved in delivery to destination cells and tissues. These functional miRNAs are supposed to use cellular machinery to regulate mRNA translation to protein once they have been delivered. Multiple sequencing systems (e.g., Illumina, Ion Torrent, and SOLiD) and bioinformatics data management skills enable in-depth miRNA sequencing (miRNA-Seq), allowing for the discovery of known and novel miRNAs [12][13], alterations [14], and possible functions [15].

2. The Role of miRNAs in Ruminants

One of the most prosperous subspecies of terrestrial mammals is the ruminant family. They live in a wide range of diverse environments around the world and have a big impact on various ecosystems, industries including agriculture, leisure activities, and cultures [16]. The ability of this group of animals to survive and procreate on low-quality, low-protein, and high-fiber plant resources is a major factor in their success. Most of the various stakeholders believe that animal health is crucial to the production of livestock; however, there is a disagreement among consumers, farmers, and veterinarians over what constitutes an acceptable health state.
The term “animal health’’ lacks a precise scientific definition (ranging from the absence of disease to a broad definition of health as a state of unrestricted physical, physiological, and psychological well-being), as well as clear standards from which the condition of animals and the quality of their feed could be properly evaluated. One of the main goals of farmers is to assure the health and performance of their livestock through the adequate provision of suitable feed. By enhancing forage quality and creating rations with more balanced ratios of forage and grain, organic farming should attempt to boost the energy content of cattle diets to improve the efficiency of protein utilization and, as a result, reduce nitrogen loss to the environment [16]. Farm animals’ cell differentiation, biological development, and physiology are all significantly influenced by miRNAs. These processes include controlling muscle growth and hypertrophy, adipose tissue expansion, oocyte maturation, and early embryonic development [17]. Recent research has demonstrated the critical roles of miRNAs in sheep [18], goats [19], and bovine [20][21] rumen development, as well as the preservation of intestinal homeostasis. As a result, the miRNA expression profiles in the rumen, small intestine (duodenum and jejunum [22]), and large intestine (cecum and colon) of sheep and cattle have been identified. The study also revealed that some miRNAs are exclusively expressed in specific intestine segments, indicating that their roles may be constrained to the local microenvironment. In addition, taxonomic differences in how miRNAs regulate gene expression typically occur during the expression and processing stages [23]. In ruminants, the colon is a vital component of the hind gut’s digestive system. Cell wall polysaccharides, cellulose, and hemicellulose are fermented and used in large part by the colon in ruminants. According to Yan and colleagues’ study [23], from a total of 1572 miRNAs discovered in the colon tissues during the analysis of colon miRNA transcriptomes in preweaning and postweaning goats, 39 differentially expressed miRNAs and 88 highly expressed miRNAs were screened, and various functions of dynamic miRNAs in the regulatory system governing colon growth in goats were discovered [23]. The study provided evidence that miRNAs have a direct role in the host’s responses to nutrition by identifying potential miRNA control mechanisms of gene expression during the switch from FD to HG feeding and phytogenic supplements [24]. The study hypothesized a potential role as a mechanism of interaction between the host and the ruminal microbiota, and suggested that this communication is bidirectional, with the microbiota influencing the host’s miRNA expression pattern and the host potentially helping to shape the gut bacterial profile through the production of specific miRNAs [25].

Potential Regulatory Functions of miRNAs in Ruminant’s Milk

The ruminant mammary gland (MG) is a crucial organ responsible for producing milk for human infant and adult nutrition [26]. Not only nutrition, genetics, breed, disease pathogens, and other environmental factors, but also post transcriptional regulation of gene expression affects MG productivity. Lactation, one of the amazing outcomes of evolution, is a very dynamic and complicated process [27]. The growth of the MG and the production and release of milk are all parts of lactation. Cattle, buffaloes, goats, sheep, and camels provide almost all the world’s milk. Yaks, horses, reindeer, and donkeys are additional less frequent milk-producing animals. Each species’ prevalence and significance vary greatly between different geographical areas and nations. Feed, water, and climate are the three main factors that influence the dairy species retained. Market demand, dietary customs, and the socioeconomic makeup of each household are other variables that could impact the existence of a dairy species (e.g., poorer families tend to rely more on small ruminants). Since milk from dairy ruminants such as cows, goats, and sheep has been shown to have positive benefits on humans, extensive work has been carried out to increase milk production and improve its nutritional value [28]. Most infant formulas are based on cow milk proteins, and both cow and goat milk are commonly used as dairy products. Results show that adults absorb significant levels of milk-derived miRNAs from commercial pasteurized milk. Additionally, investigations have shown that milk exosomes can be incorporated into kidney, intestinal, intestinal cancer, and peripheral cells, as well as into macrophages and colon cancer cells [29]. According to Golan-Gerstl and colleagues, 95 percent of the miRNAs expressed in human milk are likewise expressed in goat and bovine milk, and pasteurization of bovine milk does not appear to eliminate miRNAs [30]. Additionally, 24 validated sites that were engaged in immunomodulatory actions were shared by conserved miRNAs [31]. Unsaturated fatty acid-rich feed added to the diet can be a useful strategy to boost milk’s health-promoting qualities; albeit, the impacts on the genes and pathways involved in these processes have not yet been fully and accurately described [32]. There are a number of measures, including nutrition, seasonal feed changes, and genomic variation, that can be utilized to improve the beneficial components in milk in ruminants because the process for the synthesis of milk fat is complex and subject to multifactorial regulation [33].
Bovine, caprine, and ovine species’ genetic variations of miRNAs expressed in the mammary gland or found in milk and localized in dairy quantitative trait loci (QTLs) were examined to find variations that might be the causes of dairy features. Using whole genome data to find miRNA genetic variants expressed in the mammary gland and localized in dairy QTLs, the study identifies miRNA genetic variations of interest in the context of dairy production [34].
A. Cattle
Proteins, lipids, and amino acids, as well as other bioactive substances such as hormones and cytokines, are all readily available in cow milk, which is also a good source of many other vital nutrients. Cow milk has been commercialized and regularly used by people for growth and health benefits due to its nutritional relevance. Li and colleagues describe the miRNA expression spectra of three milk fractions (fat, whey, and cells), contrast the milk fraction miRNome profiles with those of mammary gland tissue, and determine which milk fraction miRNome profile might be a better indicator of the miRNome profile of mammary gland tissue. Their findings demonstrated that the miRNAome of mammary gland tissue was accurately represented by miRNAs from milk fat. Top expressed miRNAs in milk were functionally annotated, and this revealed their crucial regulatory roles in mammary gland functions and perhaps to milk recipients [35]. Udder diseases, particularly mastitis brought on by bacterial infections, are significant issues for the dairy industry globally. Mastitis continues to be a major issue for the dairy industry globally, resulting in significant losses every year from reduced milk production (both quantity and quality), expensive treatments, and early animal culling, as well as having a significant impact on the development of antimicrobial resistance in cattle due to the widespread use of dry cow antibiotics [36]. Improved tools that can accurately detect early mammary inflammation in cattle are urgently needed given the relevance of early disease detection for minimizing the considerable financial and animal welfare implications of mastitis globally. Monitoring mammary gland health and spotting early inflammation require different methods. The most used method is somatic cell counting (SCC), which can be carried out in large quantities of milk or as individual milk samples, directly or indirectly through colorimetric quantification, frequently with the California Mastitis Test (CMT) [36]. Mammary epithelial cells release milk fat globules through a budding mechanism that encloses a crescent of the mammary epithelial cells cytoplasm in the plasma membrane [35]. Whether the miRNAs found in milk are specific to the mammary gland or come from the blood is the key question in determining the involvement of miRNAs in lactation. Tzelos and colleagues investigated the relationship between CMT scores (0 to 3), which were derived from many individual quarter milk samples (n = 236) taken from dairy cows at various lactations, and the levels of four inflammation-associated miRNAs (bta-miR-26a, bta-miR-142-5p, bta-miR-146a, and bta-miR-223). They confirmed that higher miRNA levels during lactation 1 than later lactations were responsible for the significant lactation number effect (P 0.01) for bta-miR-26a, bta-miR-142-5p, and bta-miR-146a. They also showed that bta-miR-223 and bta-miR-142-5p levels could accurately (100% sensitivity, >81% specificity) identify early inflammation. They stated that their findings offer further evidence of the usefulness of miRNAs as early diagnostic indicators of cow mastitis [36]. Using microarray and quantitative PCR analysis, Izumi and colleagues identified variations between colostrum (100 miRNAs) and mature milk (53 known miRNAs). They confirmed that some mRNA was present in cow’s milk, but that naturally occurring miRNA and mRNA in raw milk were resistant to acidic conditions and RNase treatment. Synthesized miRNA spiked in the raw milk whey were degraded [37]. Wang’s research team used transcriptome studies of mammary gland tissues from dairy cows during the high-protein/high-fat, low-protein/low-fat, or dry periods to investigate the molecular mechanisms governing milk secretion and quality in dairy cows. They discovered 25 core differentially expressed miRNAs (DE miRNAs) that were important for mammary gland growth and epithelial cell terminal differentiation during lactation, as compared to non-lactation. Their findings suggested that during mammary gland development, DE miRNAs might function as regulators of milk quality and milk secretion [38]. Xia and colleagues used miRNA and transcriptome data from the mammary epithelial cells of dairy cattle with high (H, 4.85%) and low milk fat percentages (L, 3.41%) during mid-lactation to screen and identify differentially expressed miRNAs, candidate genes, and co-regulatory pathways related to the metabolism of milk fat. In the co-expression networks of the dairy cattle mammary system, they discovered functional miRNAs and regulatory candidate genes involved in lipid metabolism (Table 1). This information advances the understanding of potential regulatory mechanisms of genetic elements and gene signaling networks involved in milk fat metabolism [39]. Since small non-coding RNAs have been linked to various phenotypes in bovine sperm, Werry and colleagues hypothesized that some differences in bull fertility may be reflected in the levels of various miRNAs in sperm. However, efforts to identify sperm-borne molecular biomarkers of male fertility have so far failed to identify a robust profile of expressed miRNAs related to fertility. The most abundant miRNAs in both populations were identified (miRs -34b-3p, -100-5p, -191-5p, -30d-4p, and -21-5p), and variations in both the total levels and particular patterns of isomiR expression were assessed. The findings offer a thorough description of the bovine sperm miRNAome and point to numerous potential roles in fertility [40]. To support the use of milk fat globules as a source of small non-coding RNAs to diagnose mastitis, their abundance from five cows before and after lipopolysaccharide (LPS) challenge was compared. Six miRNAs that are known to be regulated in the mammary gland during inflammation were also examined. The results showed that milk fat globules might be an easily accessible source of miRNAs that are possible biomarkers to detect early mastitis, and enable the application of a quick and efficient treatment. Four miRNAs (miR-494-3p, -148a-3p, -99a-5p, and -125b-5p) were variably abundant depending on the inflammatory status [41]. The Cui study team found 497 known miRNAs and 49 new ones using mRNA sequencing in the mammary glands of milking dairy cows. One of them, miRNA-71, was expressed differently in cows whose milk contained high and low levels of protein and fat [42]. A total of 21 differentially expressed genes can be referred to as targets for some of the 71 DE miRNAs based on their prior RNA sequencing data, suggesting that they may be crucial regulators of the milk protein and fat characteristics in dairy calves [43]. The miRNAomes of five essential metabolic tissues (rumen, duodenum, jejunum, liver, and mammary gland tissues) involved in protein synthesis and metabolism from 18 dairy cows fed high- and low-quality diets were studied to better understand the molecular regulatory mechanisms of milk protein production in dairy cows. There were 340, 338, 337, 330, and 328 miRNAs expressed in the rumen, duodenum, jejunum, liver, and mammary gland tissues, respectively. The findings indicated that miRNAs expressed in these tissues may play a part in controlling the transfer of amino acid for milk production downstream, which is a critical mechanism that may be related to low milk protein under poor forage feed [44]. When the mammary gland miRNomes of two dairy cow breeds, Holstein and Montbéliarde, with different mammogenic potentials that are related to differences in dairy performance, were compared, 754 distinct miRNAs were found in the mammary glands of Holstein (n = 5) and Montbéliarde (n = 6) midlactating cows. The outcome demonstrates variations in the mammary gland miRNomes of two dairy cattle breeds, and suggests that miRNAs may have a role in the flexibility of the mammary gland and the synthesis of milk components, both of which are connected to the quantity and composition of milk [45].
With Illumina RNA-sequencing, it was discovered that miRNAs have a regulatory role in the early development of the gastrointestinal tract (GIT). As a result of the findings, which included temporal and regional variations in miRNA expression as well as a connection between miRNA expression and the microbial population in the GIT during early life, there is now additional support for the theory that host–microbial interactions regulate gut development through a different mechanism [20]. Based on the variations in meat quality attributes and 90 differently expressed mRNAs, an integrated study of miRNA and mRNA expression profiling was carried out between bulls and steers, and 18 DE miRNAs were discovered. The findings offered compelling proof of the molecular genetic regulation and gene interactions in cattle [46].
B. Goats
Goat milk is typically made into cheese in Mediterranean nations and Latin America; in Africa and South Asia, it is typically drunk raw or acidified. One of the most significant livestock animals is the goat (Caprahircus) [47]. Numerous studies on goats have looked at the impact of various feeding methods on milk fat content; nevertheless, the physiological underpinnings of this reaction are still not well understood.
A study that examined the relationship between differentially expressed miRNAs in goat mammary tissue and the fatty acid composition of goat milk found that levels of miR-183, targeting MST1 (Macrophage Stimulating 1), were positively associated with the milk’s fatty acid content [48]. The MST1 gene is targeted by miR-183 in the cytoplasm of goat mammary epithelial cells, which results in an inhibition of milk fat metabolism. One of the most crucial aspects of the nutritional quality of goat milk is its lipid composition. For instance, the findings of the study by Ollier and colleagues revealed that whole intact rapeseeds or sunflower oil in high-forage or high-concentrate diets affected milk yield and composition, but not the mammary mRNA expression of the important genes involved in lactose (for example, α-lactalbumin), protein (for example, β-casein), and lipid metabolism (e.g., lipoprotein, lipase) [49]. The response to dietary interventions did not appear to be mediated by changes in the mRNA abundance of genes encoding essential lipogenic enzymes and the associated transcription factors, according to [50] on lactating goats fed a supplement of sunflower seed oil. Thirty highly expressed miRNAs, including miR-103, whose expression correlates with lactation, were found by high-throughput sequencing in the mammary gland of lactating goats. This study’s conclusions provided new insight about the roles of miR-103 and the molecular processes that control milk fat synthesis [51]. Diets containing extruded linseed alone or in conjunction with fish oil in lactating goats exhibited effects on mRNA connected to protein metabolism and transport pathways rather than lipid metabolism, as well as a significant alteration in the FA composition of milk [52]. MiR-25 mimics in goat mammary epithelial cells lowered the expression of genes involved in lipid metabolism, which was inversely correlated with milk production at various phases of lactation. The study’s findings revealed the role of the miR-25/PGC-1beta regulatory axis during lactation and suggested that miR-25 may be involved in lipid metabolism [53] (Table 1). A total of 1487 unique miRNAs, including 45 novel miRNA candidates and 1442 known and conserved miRNAs with 378 differentially expressed and 758 co-expressed miRNAs, were found between early and late lactation. The study’s findings suggested that miRNAs may be crucial to early and late lactation throughout the development of the dairy goat mammary gland, which will help researchers better understand how genes regulate mammary gland development and lactation [54]. The goat genome was sequenced, and 487 miRNAs were identified. The greatest miRNA clusters were discovered on chromosome 21 [55]. Overall, 131 novel and 300 conserved miRNAs were identified after analyzing goat MG tissues during early lactation using the Illumina-Solexa high-throughput sequencing method [56]. Additionally, 346 conserved and 95 novel miRNAs were discovered in goat MG tissues from dry off, and peak lactation does use the same technique (Illumina-Solexa sequencing).
C. Sheep
Most sheep in the world live in developing nations. In colder regions they are even more common than goats. Sheep farming includes a variety of products that can be produced, including milk, meat, skin, fiber, and dung, although most small-scale farmers in developing nations grow sheep for meat or for sale as livestock in local markets. Milk production and lactation duration have not significantly increased because of genetic selection in dairy sheep. Awassi, East Friesian, and Lacaune are sheep breeds used for dairy products [47].
In a study with different fat-tailed sheep breeds, 155 DE miRNAs, including 78 up-regulated and 77 down-regulated miRNAs, were found between the tail fat tissue of Hu sheep (short-fat-tailed) and Tibetan sheep (short-thin-tailed) using miRNA-Seq. The findings might offer a useful theoretical framework for research into the molecular processes behind sheep tail adipogenesis [57]. Using RNA sequencing and cell-level validation (an error-based approach to design and optimization) is crucial when dealing with a complicated process such as NGS. The degree of validation and quality control required for specific process steps can be determined by carefully considering the likelihood of potential failure spots. It also helps in troubleshooting any errors and validating changes made to various test system components. The role of miRNA in the deposition of intramuscular fat (IMF) was investigated, and 59 DE-miRNAs were discovered between 2-month-old (Mth-2) and 12 month-old (Mth-12) Aohan fine-wool sheep (AFWS). In an effort to enhance the quality of sheep meat, the study identified lists of miRNAs linked to intramuscular lipid deposition in sheep and their prospective targets [58]. A study compared the microstructures and the miRNA expression profile of mammary gland (MG) tissues at peak lactation in small-tailed Han and Gansu Alpine Merino sheep, with various milk production attributes. Eighteen of the one hundred and twenty-four mature miRNAs produced were differentially expressed between the two breeds. The findings indicated that the functions of miRNAs in the growth and lactation of MG in sheep can be improved. The results also indicated that the targeted genes of differentially expressed miRNAs were mainly involved in metabolic pathways and signaling pathways related to MG development, milk protein, and fat synthesis [59] (Table 1).
Table 1. MiRNAs involved in the regulation processes of mammary lipid metabolism of ruminants.

References

  1. Cameron, A.; McAllister, T.A. Antimicrobial usage and resistance in beef production. J. Anim. Sci. Biotechnol. 2016, 7, 68.
  2. Kaur, S.; Kumar, S. Simardeep Kaur and Suresh Kumar. Nutriepigenomics: Need of the Day to Integrate Genetics, Epigenetics and Environment towards Nutritious Food for Healthy Life. Food Sci. Nutri. Technol. 2020, 5, 000239.
  3. Kumar, S.; Chinnusamy, V.; Mohapatra, T. Epigenetics of Modified DNA Bases: 5-Methylcytosine and Beyond. Front. Genet. 2018, 9, 640.
  4. Ross, S.A.; Davis, C.D. The emerging role of microRNAs and nutrition in modulating health and disease. Annu. Rev. Nutr. 2014, 34, 305–336.
  5. Filipowicz, W.; Bhattacharyya, S.N.; Sonenberg, N. Mechanisms of post-transcriptional regulation by microRNAs: Are the answers in sight? Nat. Rev. Genet. 2008, 9, 102–114.
  6. Stefani, G.; Slack, F. MicroRNAs in search of a target. Cold Spring Harb. Symp. Quant. Biol. 2006, 71, 129–134.
  7. Vella, M.C.; Slack, F.J. C. elegans microRNAs. WormBook 2005, 1–9.
  8. Thornton, J.E.; Gregory, R.I. How does Lin28 let-7 control development and disease? Trends Cell Biol. 2012, 22, 474–482.
  9. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75, 843–854.
  10. Hu, Y.; Lan, W.; Miller, D. Next-Generation Sequencing for MicroRNA Expression Profile. Methods Mol. Biol. 2017, 1617, 169–177.
  11. Matias-Garcia, P.R.; Wilson, R.; Mussack, V.; Reischl, E.; Waldenberger, M.; Gieger, C.; Anton, G.; Peters, A.; Kuehn-Steven, A. Impact of long-term storage and freeze-thawing on eight circulating microRNAs in plasma samples. PLoS ONE 2020, 15, e0227648.
  12. Mendes, N.D.; Freitas, A.T.; Sagot, M.-F. Current tools for the identification of miRNA genes and their targets. Nucleic Acids Res. 2009, 37, 2419–2433.
  13. Akhtar, M.M.; Micolucci, L.; Islam, M.S.; Olivieri, F.; Procopio, A.D. Bioinformatic tools for microRNA dissection. Nucleic Acids Res. 2016, 44, 24–44.
  14. Gardner, P.P.; Vinther, J. Mutation of miRNA target sequences during human evolution. Trends Genet. 2008, 24, 262–265.
  15. Michlewski, G.; Cáceres, J.F. Post-transcriptional control of miRNA biogenesis. RNA 2019, 25, 1–16.
  16. Mendes, R.E. Ruminants: Anatomy, Behavior, and Diseases; Nova Biomedical: Hauppauge, NY, USA, 2012.
  17. Wang, X.; Gu, Z.; Jiang, H. MicroRNAs in farm animals. Animal 2013, 7, 1567–1575.
  18. Hou, L.; Ji, Z.; Wang, G.; Wang, J.; Chao, T.; Wang, J. Identification and characterization of microRNAs in the intestinal tissues of sheep (Ovis aries). PLoS ONE 2018, 13, e0193371.
  19. Zhong, T.; Wang, C.; Hu, J.; Chen, X.; Niu, L.; Zhan, S.; Wang, L.; Guo, J.; Cao, J.; Li, L.; et al. Comparison of MicroRNA Transcriptomes Reveals the Association between MiR-148a-3p Expression and Rumen Development in Goats. Anim. Open Access J. MDPI 2020, 10, E1951.
  20. Liang, G.; Malmuthuge, N.; McFadden, T.B.; Bao, H.; Griebel, P.J.; Stothard, P.; Guan, L.L. Potential Regulatory Role of MicroRNAs in the Development of Bovine Gastrointestinal Tract during Early Life. PLoS ONE 2014, 9, e92592.
  21. Do, D.N.; Dudemaine, P.-L.; Fomenky, B.E.; Ibeagha-Awemu, E.M. Integration of miRNA weighted gene co-expression network and miRNA-mRNA co-expression analyses reveals potential regulatory functions of miRNAs in calf rumen development. Genomics 2019, 111, 849–859.
  22. Sun, H.-Z.; Chen, Y.; Guan, L.L. MicroRNA expression profiles across blood and different tissues in cattle. Sci. Data 2019, 6, 190013.
  23. Yan, Q.; Tian, L.; Chen, W.; Kang, J.; Tang, S.; Tan, Z. Developmental Alterations of Colonic microRNA Profiles Imply Potential Biological Functions in Kid Goats. Animals 2022, 12, 1533.
  24. Pacífico, C.; Ricci, S.; Sajovitz, F.; Castillo-Lopez, E.; Rivera-Chacon, R.; Petri, R.M.; Zebeli, Q.; Reisinger, N.; Kreuzer-Redmer, S. Bovine rumen epithelial miRNA-mRNA dynamics reveals post-transcriptional regulation of gene expression upon transition to high-grain feeding and phytogenic supplementation. Genomics 2022, 114, 110333.
  25. Ricci, S.; Petri, R.M.; Pacífico, C.; Castillo-Lopez, E.; Rivera-Chacon, R.; Sener-Aydemir, A.; Reisinger, N.; Zebeli, Q.; Kreuzer-Redmer, S. Characterization of presence and activity of microRNAs in the rumen of cattle hints at possible host-microbiota cross-talk mechanism. Sci. Rep. 2022, 12, 13812.
  26. Dysin, A.P.; Barkova, O.Y.; Pozovnikova, M.V. The Role of microRNAs in the Mammary Gland Development, Health, and Function of Cattle, Goats, and Sheep. Non-Coding RNA 2021, 7, 78.
  27. Do, D.N.; Ibeagha-Awemu, E.M. Non-Coding RNA Roles in Ruminant Mammary Gland Development and Lactation. In Current Topics in Lactation; Gigli, I., Ed.; InTech: London, UK, 2017; Available online: http://www.intechopen.com/books/current-topics-in-lactation/non-coding-rna-roles-in-ruminant-mammary-gland-development-and-lactation (accessed on 10 August 2022).
  28. Ibeagha-Awemu, E.M.; Li, R.; Ammah, A.A.; Dudemaine, P.-L.; Bissonnette, N.; Benchaar, C.; Zhao, X. Transcriptome adaptation of the bovine mammary gland to diets rich in unsaturated fatty acids shows greater impact of linseed oil over safflower oil on gene expression and metabolic pathways. BMC Genom. 2016, 17, 104.
  29. Baier, S.R.; Nguyen, C.; Xie, F.; Wood, J.R.; Zempleni, J. MicroRNAs are absorbed in biologically meaningful amounts from nutritionally relevant doses of cow milk and affect gene expression in peripheral blood mononuclear cells, HEK-293 kidney cell cultures, and mouse livers. J. Nutr. 2014, 144, 1495–1500.
  30. Golan-Gerstl, R.; Elbaum Shiff, Y.; Moshayoff, V.; Schecter, D.; Leshkowitz, D.; Reif, S. Characterization and biological function of milk-derived miRNAs. Mol. Nutr. Food Res. 2017, 61, 1700009.
  31. Cintio, M.; Polacchini, G.; Scarsella, E.; Montanari, T.; Stefanon, B.; Colitti, M. MicroRNA Milk Exosomes: From Cellular Regulator to Genomic Marker. Animals 2020, 10, 1126.
  32. Cremonesi, P.; Capra, E.; Turri, F.; Lazzari, B.; Chessa, S.; Battelli, G.; Colombini, S.; Rapetti, L.; Castiglioni, B. Effect of Diet Enriched With Hemp Seeds on Goat Milk Fatty Acids, Transcriptome, and miRNAs. Front. Anim. Sci. 2022, 3, 1–16.
  33. Hanuš, O.; Samková, E.; Křížová, L.; Hasoňová, L.; Kala, R. Role of Fatty Acids in Milk Fat and the Influence of Selected Factors on Their Variability—A Review. Mol. J. Synth. Chem. Nat. Prod. Chem. 2018, 23, 1636.
  34. Bourdon, C.; Boussaha, M.; Bardou, P.; Sanchez, M.-P.; Le Guillou, S.; Tribout, T.; Larroque, H.; Boichard, D.; Rupp, R.; Le Provost, F.; et al. In silico identification of variations in microRNAs with a potential impact on dairy traits using whole ruminant genome SNP datasets. Sci. Rep. 2021, 11, 19580.
  35. Li, R.; Dudemaine, P.-L.; Zhao, X.; Lei, C.; Ibeagha-Awemu, E.M. Comparative Analysis of the miRNome of Bovine Milk Fat, Whey and Cells. PLoS ONE 2016, 11, e0154129.
  36. Tzelos, T.; Ho, W.; Charmana, V.I.; Lee, S.; Donadeu, F.X. MiRNAs in milk can be used towards early prediction of mammary gland inflammation in cattle. Sci. Rep. 2022, 12, 5131.
  37. Izumi, H.; Kosaka, N.; Shimizu, T.; Sekine, K.; Ochiya, T.; Takase, M. Bovine milk contains microRNA and messenger RNA that are stable under degradative conditions. J. Dairy Sci. 2012, 95, 4831–4841.
  38. Wang, X.; Zhang, L.; Jin, J.; Xia, A.; Wang, C.; Cui, Y.; Qu, B.; Li, Q.; Sheng, C. Comparative transcriptome analysis to investigate the potential role of miRNAs in milk protein/fat quality. Sci. Rep. 2018, 8, 6250.
  39. Xia, L.; Zhao, Z.; Yu, X.; Lu, C.; Jiang, P.; Yu, H.; Li, X.; Yu, X.; Liu, J.; Fang, X.; et al. Integrative analysis of miRNAs and mRNAs revealed regulation of lipid metabolism in dairy cattle. Funct. Integr. Genom. 2021, 21, 393–404.
  40. Werry, N.; Russell, S.J.; Gillis, D.J.; Miller, S.; Hickey, K.; Larmer, S.; Lohuis, M.; Librach, C.; LaMarre, J. Characteristics of miRNAs Present in Bovine Sperm and Associations With Differences in Fertility. Front. Endocrinol. 2022, 13, 874371.
  41. Leroux, C.; Pawlowski, K.; Billa, P.-A.; Pires, J.A.A.; Faulconnier, Y. Milk fat globules as a source of microRNAs for mastitis detection. Livest. Sci. 2022, 263, 104997.
  42. Cui, X.; Zhang, S.; Zhang, Q.; Guo, X.; Wu, C.; Yao, M.; Sun, D. Comprehensive MicroRNA Expression Profile of the Mammary Gland in Lactating Dairy Cows With Extremely Different Milk Protein and Fat Percentages. Front. Genet. 2020, 11, 548268.
  43. Cui, X.; Hou, Y.; Yang, S.; Xie, Y.; Zhang, S.; Zhang, Y.; Zhang, Q.; Lu, X.; Liu, G.E.; Sun, D. Transcriptional profiling of mammary gland in Holstein cows with extremely different milk protein and fat percentage using RNA sequencing. BMC Genom. 2014, 15, 226.
  44. Wang, D.; Liang, G.; Wang, B.; Sun, H.; Liu, J.; Guan, L.L. Systematic microRNAome profiling reveals the roles of microRNAs in milk protein metabolism and quality: Insights on low-quality forage utilization. Sci. Rep. 2016, 6, 21194.
  45. Billa, P.A.; Faulconnier, Y.; Ye, T.; Chervet, M.; Le Provost, F.; Pires, J.A.A.; Leroux, C. Deep RNA-Seq reveals miRNome differences in mammary tissue of lactating Holstein and Montbéliarde cows. BMC Genom. 2019, 20, 621.
  46. Li, Y.; Yang, M.; Lou, A.; Yun, J.; Ren, C.; Li, X.; Xia, G.; Nam, K.; Yoon, D.; Jin, H.; et al. Integrated analysis of expression profiles with meat quality traits in cattle. Sci. Rep. 2022, 12, 5926.
  47. FAO. FAO Publications Catalogue 2022: April; FAO: Rome, Italy, 2022; 132p, Available online: https://www.fao.org/documents/card/en/c/cb9264en (accessed on 10 August 2022).
  48. Chen, Z.; Shi, H.; Sun, S.; Luo, J.; Zhang, W.; Hou, Y.; Loor, J.J. MiR-183 regulates milk fat metabolism via MST1 in goat mammary epithelial cells. Gene 2018, 646, 12–19.
  49. Ollier, S.; Leroux, C.; de la Foye, A.; Bernard, L.; Rouel, J.; Chilliard, Y. Whole intact rapeseeds or sunflower oil in high-forage or high-concentrate diets affects milk yield, milk composition, and mammary gene expression profile in goats. J. Dairy Sci. 2009, 92, 5544–5560.
  50. Toral, P.G.; Bernard, L.; Delavaud, C.; Gruffat, D.; Leroux, C.; Chilliard, Y. Effects of fish oil and additional starch on tissue fatty acid profile and lipogenic gene mRNA abundance in lactating goats fed a diet containing sunflower-seed oil. Anim. Int. J. Anim. Biosci. 2013, 7, 948–956.
  51. Lin, X.; Luo, J.; Zhang, L.; Wang, W.; Gou, D. MiR-103 Controls Milk Fat Accumulation in Goat (Capra hircus) Mammary Gland during Lactation. PLoS ONE 2013, 8, e79258.
  52. Faulconnier, Y.; Bernard, L.; Boby, C.; Domagalski, J.; Chilliard, Y.; Leroux, C. Extruded linseed alone or in combination with fish oil modifies mammary gene expression profiles in lactating goats. Anim. Int. J. Anim. Biosci. 2018, 12, 1564–1575.
  53. Ma, L.; Qiu, H.; Chen, Z.; Li, L.; Zeng, Y.; Luo, J.; Gou, D. miR-25 modulates triacylglycerol and lipid accumulation in goat mammary epithelial cells by repressing PGC-1beta. J. Anim. Sci. Biotechnol. 2018, 9, 48.
  54. Ji, Z.; Liu, Z.; Chao, T.; Hou, L.; Fan, R.; He, R.; Wang, G.; Wang, J. Screening of miRNA profiles and construction of regulation networks in early and late lactation of dairy goat mammary glands. Sci. Rep. 2017, 7, 11933.
  55. Dong, Y.; Xie, M.; Jiang, Y.; Xiao, N.; Du, X.; Zhang, W.; Tosser-Klopp, G.; Wang, J.; Yang, S.; Liang, J.; et al. Sequencing and automated whole-genome optical mapping of the genome of a domestic goat (Capra hircus). Nat. Biotechnol. 2013, 31, 135–141.
  56. Ji, Z.; Wang, G.; Xie, Z.; Zhang, C.; Wang, J. Identification and characterization of microRNA in the dairy goat (Capra hircus) mammary gland by Solexa deep-sequencing technology. Mol. Biol. Rep. 2012, 39, 9361–9371.
  57. Fei, X.; Jin, M.; Wang, Y.; Li, T.; Lu, Z.; Yuan, Z.; Wang, H.; Lu, J.; Quan, K.; Di, R.; et al. Transcriptome reveals key microRNAs involved in fat deposition between different tail sheep breeds. PLoS ONE 2022, 17, e0264804.
  58. Han, F.; Zhou, L.; Zhao, L.; Wang, L.; Liu, L.; Li, H.; Qiu, J.; He, J.; Liu, N. Identification of miRNA in Sheep Intramuscular Fat and the Role of miR-193a-5p in Proliferation and Differentiation of 3T3-L1. Front. Genet. 2021, 12, 633295.
  59. Hao, Z.Y.; Wang, J.Q.; Luo, Y.L.; Liu, X.; Li, S.B.; Zhao, M.L.; Jin, X.Y.; Shen, J.Y.; Ke, N.; Song, Y.Z.; et al. Deep small RNA-Seq reveals microRNAs expression profiles in lactating mammary gland of 2 sheep breeds with different milk performance. Domest. Anim. Endocrinol. 2021, 74, 106561.
  60. Cai, X.; Liu, Q.; Zhang, X.; Ren, Y.; Lei, X.; Li, S.; Chen, Q.; Deng, K.; Wang, P.; Zhang, H.; et al. Identification and analysis of the expression of microRNA from lactating and nonlactating mammary glands of the Chinese swamp buffalo. J. Dairy Sci. 2017, 100, 1971–1986.
  61. Shen, B.; Pan, Q.; Yang, Y.; Gao, Y.; Liu, X.; Li, W.; Han, Y.; Yuan, X.; Qu, Y.; Zhao, Z. miR-224 Affects Mammary Epithelial Cell Apoptosis and Triglyceride Production by Downregulating ACADM and ALDH2 Genes. DNA Cell Biol. 2017, 36, 26–33.
  62. Chu, M.; Zhao, Y.; Yu, S.; Hao, Y.; Zhang, P.; Feng, Y.; Zhang, H.; Ma, D.; Liu, J.; Cheng, M.; et al. MicroRNA-221 may be involved in lipid metabolism in mammary epithelial cells. Int. J. Biochem. Cell Biol. 2018, 97, 118–127.
  63. Qiaoqiao, C.; Li, H.; Liu, X.; Yan, Z.; Zhao, M.; Xu, Z.; Wang, Z.; Shi, K. MiR-24-3p regulates cell proliferation and milk protein synthesis of mammary epithelial cells through menin in dairy cows. J. Cell. Physiol. 2019, 234, 1522–1533.
  64. Wang, H.; Luo, J.; Chen, Z.; Cao, W.T.; Xu, H.F.; Gou, D.M.; Zhu, J.J. MicroRNA-24 can control triacylglycerol synthesis in goat mammary epithelial cells by targeting the fatty acid synthase gene. J. Dairy Sci. 2015, 98, 9001–9014.
  65. Li, D.; Xie, X.; Wang, J.; Bian, Y.; Li, Q.; Gao, X.; Wang, C. MiR-486 regulates lactation and targets the PTEN gene in cow mammary glands. PLoS ONE 2015, 10, e0118284.
  66. Shen, B.; Yang, Z.; Han, S.; Zou, Z.; Liu, J.; Nie, L.; Dong, W.; Li, E.; Liu, S.; Zhao, Z.; et al. Bta-miR-124a Affects Lipid Metabolism by Regulating PECR Gene. BioMed Res. Int. 2019, 2019, 2596914.
  67. Ji, Z.; Dong, F.; Wang, G.; Hou, L.; Liu, Z.; Chao, T.; Wang, J. miR-135a Targets and Regulates Prolactin Receptor Gene in Goat Mammary Epithelial Cells. DNA Cell Biol. 2015, 34, 534–540.
  68. Chen, Z.; Chu, S.; Wang, X.; Fan, Y.; Zhan, T.; Arbab, A.A.I.; Li, M.; Zhang, H.; Mao, Y.; Loor, J.J.; et al. MicroRNA-106b Regulates Milk Fat Metabolism via ATP Binding Cassette Subfamily A Member 1 (ABCA1) in Bovine Mammary Epithelial Cells. J. Agric. Food Chem. 2019, 67, 3981–3990.
  69. Wang, H.; Shi, H.; Luo, J.; Yi, Y.; Yao, D.; Zhang, X.; Ma, G.; Loor, J.J. MiR-145 Regulates Lipogenesis in Goat Mammary Cells Via Targeting INSIG1 and Epigenetic Regulation of Lipid-Related Genes. J. Cell. Physiol. 2017, 232, 1030–1040.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 323
Revisions: 2 times (View History)
Update Date: 20 Feb 2023
1000/1000
ScholarVision Creations